Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 22
Filter
Add more filters










Publication year range
1.
Langmuir ; 25(10): 5496-503, 2009 May 19.
Article in English | MEDLINE | ID: mdl-19388632

ABSTRACT

The interiors of reverse micelles formed using nonionic surfactants to sequester water droplets in a nonpolar environment have been investigated using the decavanadate molecule as a probe. Chemical shifts and line widths of the three characteristic signals in the 51V NMR spectrum of decavanadate, corresponding to vanadium atoms in equatorial peripheral, equatorial interior, and axial locations, measure the local proton concentration and characteristics of the reverse micellar interior near the decavandate probe. All samples investigated indicate deprotonation of the vanadate probe in the reverse micelle environment. However, the relative mobility of the decavanadate molecule depends on the reverse micellar components. Specifically, the 51V NMR signals of the decavandate in reverse micelles formed using only the Igepal CO-520 surfactant display sharp signals indicating that the decavandate molecule tumbles relatively freely while reverse micelles formed from a mixture of Igepal CO-610 and -430 present a more viscous environment for the decavanadate molecule; the nature of the interior of the nonionic reverse water pool varies significantly depending on the specific Igepal. The 51V NMR spectra also indicate that the interior core water pool of the reverse micelles is less acidic than the bulk aqueous solution from which the samples were created. Together, these data provide a description that allows for a comparison of the water pools in these different nonionic reverse micelles.

2.
J Inorg Biochem ; 85(1): 33-42, 2001 May.
Article in English | MEDLINE | ID: mdl-11377693

ABSTRACT

Vanadyl sulfate (VOSO(4)) was given orally to 16 subjects with type 2 diabetes mellitus for 6 weeks at a dose of 25, 50, or 100 mg vanadium (V) daily [Goldfine et al., Metabolism 49 (2000) 1-12]. Elemental V was determined by graphite furnace atomic absorption spectrometry (GFAAS). There was no correlation of V in serum with clinical response, determined by reduction of mean fasting blood glucose or increased insulin sensitivity during euglycemic clamp. To investigate the effect of administering a coordinated V, plasma glucose levels were determined in streptozotocin (STZ)-induced diabetic rats treated with the salt (VOSO(4)) or the coordinated V compound bis(maltolato)oxovandium(IV) (abbreviated as VO(malto)(2)) administered by intraperitoneal (i.p.) injection. There was no relationship of blood V concentration with plasma glucose levels in the animals treated with VOSO(4), similar to our human diabetic patients. However, with VO(malto)(2) treatment, animals with low plasma glucose tended to have high blood V. To determine if V binding to serum proteins could diminish biologically active serum V, binding of both VOSO(4) and VO(malto)(2) to human serum albumin (HSA), human apoTransferrin (apoHTf) and pig immunoglobulin (IgG) was studied with EPR spectroscopy. Both VOSO(4) and VO(malto)(2) bound to HSA and apoHTf forming different V-protein complexes, while neither V compound bound to the IgG. VOSO(4) and VO(malto)(2) showed differences when levels of plasma glucose and blood V in diabetic rodents were compared, and in the formation of V-protein complexes with abundant serum proteins. These data suggest that binding of V compounds to ligands in blood, such as proteins, may affect the available pool of V for biological effects.


Subject(s)
Diabetes Mellitus, Type 2/blood , Diabetes Mellitus, Type 2/drug therapy , Hypoglycemic Agents/pharmacology , Pyrones/pharmacology , Vanadates/pharmacology , Vanadium Compounds/pharmacology , Animals , Apoproteins/chemistry , Apoproteins/metabolism , Biological Availability , Blood Glucose/analysis , Blood Glucose/metabolism , Diabetes Mellitus, Experimental/blood , Diabetes Mellitus, Experimental/drug therapy , Electron Spin Resonance Spectroscopy , Fasting , Humans , Hypoglycemic Agents/chemistry , Hypoglycemic Agents/metabolism , Immunoglobulin G/chemistry , Immunoglobulin G/metabolism , Male , Pyrones/chemistry , Pyrones/metabolism , Rats , Rats, Wistar , Serum Albumin/chemistry , Serum Albumin/metabolism , Streptozocin , Transferrin/chemistry , Transferrin/metabolism , Treatment Outcome , Vanadates/chemistry , Vanadates/metabolism , Vanadium/blood , Vanadium/urine , Vanadium Compounds/chemistry , Vanadium Compounds/metabolism
3.
J Biol Inorg Chem ; 6(1): 82-90, 2001 Jan.
Article in English | MEDLINE | ID: mdl-11191225

ABSTRACT

Eight scorpionate-zinc thiolate complexes, [(L1O)ZnSPh], [(L1O)ZnSPhF5], [(L1O)ZnSBz], [(L1O)ZnSPh2,6-Me], [(L1O)ZnSPh2,4-Me], [(L1O)ZnSPh4-NO2], [(TpPh,Ph)ZnSPh], and [(L2S)ZnSPh], were reacted with methyl iodide in chloroform, liberating the corresponding methyl thioethers as determined by 1H NMR. Three of these complexes are new and their synthesis and structural characterization are reported here. Weak alkylating agents such as trimethyl phosphate failed to undergo methyl transfer to the zinc thiolates under these conditions. Analysis of kinetic data as a function of concentration, temperature, pKa of the exogenous thiolate, and donor atom of the tripodal ligand are consistent with a mechanism where the zinc-bound thiolate is the active nucleophile in an associative-type methyl transfer reaction. Our model studies also provide experimental evidence to support the hypothesis that some enzymes can use the charge of the metal coordination site to modulate catalytic activity.


Subject(s)
Enzymes/chemistry , Models, Chemical , Organometallic Compounds/chemistry , Sulfur/chemistry , Zinc/chemistry , Crystallography, X-Ray , Methylation
4.
J Inorg Biochem ; 80(1-2): 123-31, 2000 May 30.
Article in English | MEDLINE | ID: mdl-10885472

ABSTRACT

The chemistry of vanadium compounds that can be taken orally is very timely since a vanadium(IV) compound, KP-102, is currently in clinical trials in humans, and the fact that human studies with inorganic salts have recently been reported. VO(acac)2 and VO(Et-acac)2 (where acac is acetylacetonato and Et-acac is 3-ethyl-2,4-pentanedionato) have long-term in vivo insulin mimetic effects in streptozotocin induced diabetic Wistar rats. Structural characterization of VO(acac)2 and two derivatives, VO(Me-acac)2 and VO(Et-acac)2, in the solid state and solution have begun to delineate the size limits of the insulin-like active species. Oral ammonium dipicolinatooxovanadium(V) is a clinically useful hypoglycemic agent in cats with naturally occurring diabetes mellitus. This compound is particularly interesting since it represents the first time that a well-characterized organic vanadium compound with the vanadium in oxidation state five has been found to be an orally effective hypoglycemic agent in animals.


Subject(s)
Hypoglycemic Agents/chemistry , Insulin/pharmacology , Vanadium Compounds/chemistry , Vanadium/chemistry , Administration, Oral , Animals , Blood Glucose/metabolism , Cats , Diabetes Mellitus, Experimental/blood , Diabetes Mellitus, Experimental/drug therapy , Humans , Hydrogen-Ion Concentration , Hypoglycemic Agents/pharmacology , Magnetic Resonance Spectroscopy , Molecular Structure , Rats , Rats, Wistar , Stereoisomerism , Vanadium/pharmacology , Vanadium Compounds/pharmacology
5.
Inorg Chem ; 39(3): 406-16, 2000 Feb 07.
Article in English | MEDLINE | ID: mdl-11229556

ABSTRACT

The syntheses and the solid state structural and spectroscopic solution characterizations of VO(Me-acac)2 and VO(Et-acac)2 (where Me-acac is 3-methyl-2,4-pentanedionato and Et-acac is 3-ethyl-2,4-pentanedionato) have been conducted since both VO(acac)2 and VO(Et-acac)2 have long-term in vivo insulin-mimetic effects in streptozotocin-induced diabetic Wistar rats. X-ray structural characterizations of VO(Me-acac)2 and VO(Et-acac)2 show that both contain five-coordinate vanadium similar to the parent VO(acac)2. The unit cells for VO(Et-acac)2 and VO(Me-acac)2 are both triclinic, P1, with a = 9.29970(10) A, b = 13.6117(2) A, c = 13.6642(2) A, alpha = 94.1770(10) degrees, beta = 106.4770(10) degrees, gamma = 106.6350(10) degrees for VO(Et-acac)2 and a = 7.72969(4) A, b = 8.1856(5) A, c = 11.9029(6) A, alpha = 79.927(2) degrees, beta = 73.988(2)degrees, gamma = 65.1790(10)degrees for VO(Me-acac)2. The total concentration of EPR-observable vanadium(IV) species for VO(acac)2 and derivatives in water solution at 20 degreesC was determined by double integration of the EPR spectra and apportioned between individual species on the basis of computer simulations of the spectra. Three species were observed, and the concentrations were found to be time, pH, temperature, and salt dependent. The three complexes are assigned as the trans-VO(acac)2.H2O adduct, cis-VO(acac)2.H2O adduct, and a hydrolysis product containing one vanadium atom and one R-acac- group. The reaction rate for conversion of species was slower for VO(acac)2 than for VO(malto)2, VO(Et-acac)2, and VO(Me-acac)2; however, in aqueous solution the rates for all of these species are slow compared to those of other vanadium species. The concentration of vanadium(V) species was determined by 51V NMR. The visible spectra were time dependent, consistent with the changes in species concentrations that were observed in the EPR and NMR spectra. EPR and visible spectroscopic studies of solutions prepared as for administration to diabetic rats documented both a salt effect on speciation and formation of a new halogen-containing complex. Compound efficacy with respect to long-term lowering of plasma glucose levels in diabetic rats traces the concentration of the hydrolysis product in the administration solution.


Subject(s)
Insulin/physiology , Organometallic Compounds/chemistry , Vanadates/chemistry , Animals , Blood Glucose/drug effects , Crystallography, X-Ray , Diabetes Mellitus, Experimental/chemically induced , Disease Models, Animal , Electron Spin Resonance Spectroscopy , Insulin/chemistry , Magnetic Resonance Spectroscopy , Male , Molecular Mimicry , Molecular Structure , Organometallic Compounds/pharmacology , Organometallic Compounds/therapeutic use , Rats , Rats, Wistar , Streptozocin , Structure-Activity Relationship , Vanadates/pharmacology , Vanadates/therapeutic use
6.
Br J Pharmacol ; 126(2): 467-77, 1999 Jan.
Article in English | MEDLINE | ID: mdl-10077240

ABSTRACT

1. Vanadium compounds can mimic actions of insulin through alternative signalling pathways. The effects of three organic vanadium compounds were studied in non-ketotic, streptozotocin-diabetic rats: vanadyl acetylacetonate (VAc), vanadyl 3-ethylacetylacetonate (VEt), and bis(maltolato)oxovanadium (VM). A simple inorganic vanadium salt, vanadyl sulphate (VS) was also studied. 2. Oral administration of the three organic vanadium compounds (125 mg vanadium element 1(-1) in drinking fluids) for up to 3 months induced a faster and larger fall in glycemia (VAc being the most potent) than VS. Glucosuria and tolerance to a glucose load were improved accordingly. 3. Activities and mRNA levels of key glycolytic enzymes (glucokinase and L-type pyruvate kinase) which are suppressed in the diabetic liver, were restored by vanadium treatment. The organic forms showed greater efficacy than VS, especially VAc. 4. VAc rats exhibited the highest levels of plasma or tissue vanadium, most likely due to a greater intestinal absorption. However, VAc retained its potency when given as a single i.p. injection to diabetic rats. Moreover, there was no relationship between plasma or tissue vanadium levels and any parameters of glucose homeostasis and hepatic glucose metabolism. Thus, these data suggest that differences in potency between compounds are due to differences in their insulin-like properties. 5. There was no marked toxicity observed on hepatic or renal function. However, diarrhoea occurred in 50% of rats chronically treated with VS, but not in those receiving the organic compounds. 6. In conclusion, organic vanadium compounds, in particular VAc, correct the hyperglycemia and impaired hepatic glycolysis of diabetic rats more safely and potently than VS. This is not simply due to improved intestinal absorption, indicating more potent insulin-like properties.


Subject(s)
Glucose/metabolism , Ligands , Vanadium Compounds/pharmacology , Administration, Oral , Animals , Blood Glucose/drug effects , Blood Glucose/metabolism , Body Weight/drug effects , Diabetes Mellitus, Experimental/metabolism , Disinfectants/pharmacology , Glucokinase/drug effects , Glucokinase/genetics , Glucokinase/metabolism , Hydroxybutyrates/chemistry , Hydroxybutyrates/pharmacology , Hypoglycemic Agents/pharmacology , Injections, Intraperitoneal , Insulin/metabolism , Islets of Langerhans/drug effects , Islets of Langerhans/metabolism , Liver/drug effects , Liver/metabolism , Liver Glycogen/metabolism , Male , Muscles/drug effects , Muscles/metabolism , Organometallic Compounds/pharmacology , Pentanones/chemistry , Pentanones/pharmacology , Phosphoenolpyruvate Carboxykinase (GTP)/drug effects , Phosphoenolpyruvate Carboxykinase (GTP)/genetics , Phosphoenolpyruvate Carboxykinase (GTP)/metabolism , Pyrones/chemistry , Pyrones/pharmacology , Pyruvate Kinase/drug effects , Pyruvate Kinase/genetics , Pyruvate Kinase/metabolism , RNA, Messenger/drug effects , RNA, Messenger/genetics , Rats , Rats, Wistar , Time Factors , Vanadates/chemistry , Vanadates/pharmacology , Vanadium Compounds/chemistry
7.
Arch Biochem Biophys ; 338(1): 7-14, 1997 Feb 01.
Article in English | MEDLINE | ID: mdl-9015381

ABSTRACT

Vanadium compounds mimic insulin actions in different cell types. The present study concerns the insulin-like effects of three vanadium(V) derivatives and one vanadium(IV) complex on osteoblast-like (UMR106 and MC3T3E1) cells in culture. The vanadium oxalate and vanadium citrate complexes hydrolyzed completely under the culture conditions, whereas more than 40% of the vanadium tartrate and nitrilotriacetate complexes remained. Vanadate, as well as vanadium oxalate, citrate, and tartrate complexes enhanced cell proliferation (as measured by the crystal violet assay), glucose consumption, and protein content in UMR106 and MC3T3E1 osteoblast-like cells. The vanadium nitrilotriacetate complex (the only peroxo complex tested) stimulated cell proliferation in UMR106 but not in MC3T3E1 cells. This derivative strongly transformed the morphology of the MC3T3E1 cells. All vanadium(V) compounds inhibited cell differentiation (alkaline phosphatase activity) in UMR106 cells. Our data are consistent with the interpretation that vanadium oxalate and citrate complexes hydrolyze to vanadate. Vanadium nitrilotriacetate would appear to be toxic for normal MC3T3E1 osteoblasts. In contrast, the vanadium tartrate complex induced a proliferative effect; however, it did not alter cell differentiation.


Subject(s)
Insulin/pharmacology , Osteoblasts/drug effects , Vanadium Compounds/pharmacology , 3T3 Cells , Animals , Cell Differentiation/drug effects , Cell Division/drug effects , Cell Line , Culture Media , Drug Stability , Glucose/metabolism , Magnetic Resonance Spectroscopy , Mice , Osteoblasts/cytology , Osteoblasts/metabolism , Proteins/metabolism , Vanadates/pharmacology , Vanadium Compounds/chemistry , Vanadium Compounds/metabolism
8.
Biochem J ; 321 ( Pt 2): 333-9, 1997 Jan 15.
Article in English | MEDLINE | ID: mdl-9020863

ABSTRACT

Vanadium oxoions have been shown to elicit a wide range of effects in biological systems, including an increase in the quantity of phosphorylated proteins. This response has been attributed to the inhibition of protein phosphatases, the indirect activation of protein kinases via stimulation of enzymes at early steps in signal transduction pathways and/or the direct activation of protein kinases. We have evaluated the latter possibility by exploring the effects of vanadate, decavanadate and vanadyl cation species on the activity of the cAMP-dependent protein kinase (PKA), a serine/threonine kinase. Vanadate, in the form of monomer, dimer, tetramer and pentamer species, neither inhibits nor activates PKA. In marked contrast, decavandate is a competitive inhibitor (Ki = 1.8 +/- 0.1 mM) of kemptide (Leu-Arg-Arg-Ala-Ser-Leu-Gly), a peptide-based substrate. This inhibition pattern is especially surprising, since the negatively charged decavanadate would not be predicted to bind to the region of the active site of the enzyme that accommodates the positively charged kemptide substrate. Our studies suggest that decavanadate can associate with kemptide in solution, which would prevent kemptide from interacting with the enzyme. Vanadium(IV) also inhibits the PKA-catalysed phosphorylation of kemptide, but with an IC50 of 366 +/- 10 microM. However, in this case V4+ appears to bind to the Mg(2+)-binding site, since it can substitute for Mg2+. In the absence of Mg2+, the optimal concentration of vanadium(IV) for the PKA-catalysed phosphorylation of kemptide is 100 microM, with concentrations above 100 microM being markedly inhibitory. However, even at the optimal 100 microM V4+ concentration, the Vmax and K(m) values (for kemptide) are significantly less favourable than those obtained in the presence of 100 microM Mg2+. In summary, we have found that oxovanadium ions can directly alter the activity of the serine/threonine-specific PKA.


Subject(s)
Cyclic AMP-Dependent Protein Kinases/antagonists & inhibitors , Vanadates/pharmacology , Animals , Binding, Competitive , Catalysis , Cations , Cattle , Cyclic AMP-Dependent Protein Kinases/metabolism , Oligopeptides/metabolism , Spectrophotometry, Ultraviolet , Substrate Specificity/drug effects , Vanadates/metabolism
9.
Biochemistry ; 35(25): 8314-8, 1996 Jun 25.
Article in English | MEDLINE | ID: mdl-8679588

ABSTRACT

Both exogenously added vanadate (oxidation state +5) and vanadyl (oxidation state +4) mimic the rapid responses of insulin through alternative signaling pathways, not involving insulin receptor activation [reviewed in Shechter et al. (1995) Mol. Cell. Biochem. 153, 39-47]. Vanadium exhibits complex chemistry, fluctuating between vanadate(+5) and vanadyl(+4), according to the prevailing conditions. Using several experimental approaches, we report here on a distinct vanadate(+5)-independent, vanadyl(+4)-dependent activating pathway. The key components of this pathway are membrane protein phosphotyrosine phosphatases (PTPases) and a cytosolic (nonreceptor) protein-tyrosine kinase (CytPTK). We further suggest that vanadate(+5) is not reduced rapidly to vanadyl(+4) inside the cell, and entered vanadyl sulfate(+4) is capable of undergoing spontaneous oxidation to vanadate(+5) in vivo. Finally, we show that the promotion and full expression of a downstream bioeffect such as lipogenesis requires both activation of CytPTK and prolonged stability of vanadyl(+4) against oxidation.


Subject(s)
Adipocytes/drug effects , Insulin/pharmacology , Vanadium Compounds/pharmacology , Animals , Cytosol/enzymology , Dose-Response Relationship, Drug , Enzyme Activation , Glutathione/pharmacology , Hypoglycemic Agents/pharmacology , Lipids/biosynthesis , Male , Molecular Mimicry , Oxidation-Reduction , Protein-Tyrosine Kinases/metabolism , Rats , Rats, Wistar , Signal Transduction
10.
Mol Cell Biochem ; 153(1-2): 17-24, 1995.
Article in English | MEDLINE | ID: mdl-8927035

ABSTRACT

The stability of 11 vanadium compounds is tested under physiological conditions and in administration fluids. Several compounds including those currently used as insulin-mimetic agents in animal and human studies are stable upon dissolution in distilled water but lack such stability in distilled water at pH 7. Complex lability may result in decomposition at neutral pH and thus may compromise the effectiveness of these compounds as therapeutic agents; Even well characterized vanadium compounds are surprisingly labile. Sufficiently stable complexes such as the VEDTA complex will only slowly reduce, however, none of the vanadium compounds currently used as insulin-mimetic agents show the high stability of the VEDTA complex. Both the bis(maltolato)oxovanadium(IV) and peroxovanadium complexes extend the insulin-mimetic action of vanadate in reducing cellular environments probably by increased lifetimes under physiological conditions and/or by decomposing to other insulin mimetic compounds. For example, treatment with two equivalents of glutathione or other thiols the (dipicolinato)peroxovanadate(V) forms (dipicolinato)oxovanadate(V) and vanadate, which are both insulin-mimetic vanadium(V) compounds and can continue to act. The reactivity of vanadate under physiological conditions effects a multitude of biological responses. Other vanadium complexes may mimic insulin but not induce similar responses if the vanadate formation is blocked or reduced. We conclude that three properties, stability, lability and redox chemistry are critical to prolong the half-life of the insulin-mimetic form of vanadium compounds under physiological conditions and should all be considered in development of vanadium-based oral insulin-mimetic agents.


Subject(s)
Hypoglycemic Agents/chemistry , Vanadium Compounds/chemistry , Vanadium/chemistry , Hypoglycemic Agents/metabolism , Oxidation-Reduction , Vanadium Compounds/metabolism
12.
Met Ions Biol Syst ; 31: 287-324, 1995.
Article in English | MEDLINE | ID: mdl-8564811

ABSTRACT

The chemical similarities between vanadate and phosphate combined with the ability of vanadate to readily undergo changes in coordination geometry allows this ion to strongly influence the function of a large variety of phosphate-metabolizing enzymes. As transition state analogs, spontaneously formed vanadate complexes are potent inhibitors of a number of enzymes, including some ribonucleases, mutases, and phosphatases. In addition, vanadate is an effective inhibitor of many ATPases, kinases, lyases, and synthases. Vanadate oligomers tend to be weaker inhibitors than vanadate but do influence the function of dehydrogenases, mutases, aldolases, kinases, and others. Of the oligomers, decavanadate is unique in that it seems to bind only in polyphosphate binding domains. Peroxovanadate has not yet been well studied but it seems to inhibit enzymes that do not utilize a pentacoordinate vanadate in the catalysis cycle. Additional detailed studies of vanadate-initiated inhibition of enzymes will expand our understanding of the various mechanisms of action of vanadate and its derivatives that have been briefly described here and will doubtless provide insight into other functions of this unique material.


Subject(s)
Enzyme Inhibitors/pharmacology , Phosphates/metabolism , Phosphoric Monoester Hydrolases/antagonists & inhibitors , Phosphotransferases/antagonists & inhibitors , Vanadates/pharmacology , Adenosine Triphosphatases/antagonists & inhibitors , Animals , Binding Sites/drug effects , Enzyme Inhibitors/chemistry , Lyases/antagonists & inhibitors , Ribonucleases/antagonists & inhibitors , Substrate Specificity , Vanadates/chemistry
13.
Anal Biochem ; 209(1): 85-94, 1993 Feb 15.
Article in English | MEDLINE | ID: mdl-8465966

ABSTRACT

31P NMR spectroscopy was used to compare the phosphorus compound content in aqueous, acidic, and organic extracts of Phaseolus vulgaris seeds. Two chloroform-methanol and three ethanol extractions were used to isolate phospholipids from dry cotyledons and the phospholipid profiles were compared. Variations in phospholipid composition and artificially high concentrations of lysophosphatidylcholine were observed in the ethanol extracts. Other phosphorus compounds were extracted using perchloric acid, trichloroacetic acid, trichloroacetic acid in ether, hydrochloric acid, boiling water, and aqueous Hepes. The Hepes extraction was introduced in order to compare the results of a gentle procedure with those involving heat and acid treatments. Low concentrations of phospho-sugars and other phosphorus metabolites were found in all these extracts. High concentrations of phytate were found in all aqueous and acidic extractions; however, 31P NMR spectra of aqueous extractions did not show the phytate resonance. The NMR silence of the phytate resonances is attributed to complexation of phytate. The 31P NMR spectra of aqueous extracts contained a broad resonance at 0.1 ppm assigned to protein-bound RNA. Various extraction procedures on the same biological material are presented, and the comparison of these methods will facilitate future analysis and interpretation of literature reports involving these methods of tissue analysis.


Subject(s)
Fabaceae/chemistry , Phosphorus/metabolism , Plants, Medicinal , Seeds/chemistry , Chemistry Techniques, Analytical/methods , Chloroform/chemistry , Fabaceae/metabolism , HEPES/chemistry , Hydrogen-Ion Concentration , Magnetic Resonance Spectroscopy/methods , Methanol/chemistry , Phospholipids/analysis , Phospholipids/metabolism , Phosphorus/analysis , Plant Extracts/analysis , Plant Extracts/metabolism , Plant Proteins/analysis , Plant Proteins/metabolism , Seeds/metabolism , Water/chemistry
14.
Biochemistry ; 31(47): 11731-9, 1992 Dec 01.
Article in English | MEDLINE | ID: mdl-1332769

ABSTRACT

Uteroferrin, the purple acid phosphatase from porcine uterine fluid, is noncompetitively inhibited by vanadate in a time-dependent manner under both aerobic and anaerobic conditions. This time-dependent inhibition is observed only with the diiron enzyme and is absent when the FeZn enzyme is used. The observations are attributed to the sequential formation of two uteroferrin-vanadium complexes. The first complex forms rapidly and reversibly, while the second complex forms slowly and results in the production of catalytically inactive oxidized uteroferrin and V(IV), which is observed by EPR. The redox reaction can be reversed by treatment of the oxidized enzyme first with (V(IV)) and then EDTA to generate a catalytically active uteroferrin. Multiple inhibition kinetics suggests that vanadate is mutually exclusive with molybdate, tungstate, and vanadyl cation. The binding site for each of these anions is distinct from the site to which the competitive inhibitors phosphate and arsenate bind. The time-dependent inhibition by vanadate of uteroferrin containing the diiron core represents a new type of mechanism by which vanadium can interact with proteins and gives additional insight into the binding of anions to uteroferrin.


Subject(s)
Metalloproteins/antagonists & inhibitors , Tungsten Compounds , Vanadates/pharmacology , Acid Phosphatase , Animals , Anions , Arsenates/pharmacology , Binding Sites , Cations , Edetic Acid/pharmacology , Electron Spin Resonance Spectroscopy , Female , Iron/analysis , Isoenzymes , Kinetics , Metalloproteins/chemistry , Metalloproteins/metabolism , Molybdenum/pharmacology , Oxidation-Reduction , Phosphates/pharmacology , Rhenium/pharmacology , Swine , Tartrate-Resistant Acid Phosphatase , Tungsten/pharmacology , Vanadates/metabolism , Vanadium/metabolism
15.
Biochemistry ; 31(29): 6812-21, 1992 Jul 28.
Article in English | MEDLINE | ID: mdl-1637817

ABSTRACT

Reductive, nonreductive, and photolytic interactions of vanadate with fructose-1,6-bisphosphate aldolase were examined and used to explore the interactions of oxoanions with aldolase. Aldolase is known to interact strongly with oxoanions at low ionic strength and weakly at higher ionic strength. Oxoanions inhibit aldolase competitively with respect to fructose 1,6-bisphosphate although the location of the oxoanion binding site on aldolase remains elusive. In this work, the interaction of aldolase with a series of oxoanions was compared at ionic strength approaching physiologic levels. The size and shape of the anion were important for the effective binding to aldolase, and no significant increase in affinity for aldolase was observed by the addition of alkyl groups to the oxoanions. Vanadate competitively inhibits aldolase in a manner analogous to the other oxoanions. Since vanadate solutions contain a mixture of vanadate oxoanions, the nature of the inhibition was determined using a combination of enzyme kinetics and 51V NMR spectroscopy. Aldolase contains a significant number of thiol functionalities, and as expected, vanadate undergoes redox chemistry with them, generating an irreversibly inhibited aldolase. This oxidative chemistry was attributed to the vanadate tetramer, whereas vanadate dimer was a reversible inhibitor. Vanadate monomer does not significantly interact with aldolase reversibly or irreversibly. Vanadyl cation has the lowest inhibition constant under these high ionic strength conditions. Using Yonetani-Theorell analysis, it appears that phosphate, pyrophosphate, and sulfate bind to the same site on aldolase, whereas vanadate, arsenate, and molybdate bind to another site. UV light-induced photocleavage of aldolase by vanadate was examined, and the loss of aldolase activity was correlated with cleavage of the aldolase subunit. Further studies using vanadium as a probe should reveal details on the location of the vanadate and vanadyl cation binding sites. This study suggests several sites on aldolase will accommodate oxoanions, and one of these sites also accommodates vanadyl cation.


Subject(s)
Fructose-Bisphosphate Aldolase/metabolism , Muscles/enzymology , Vanadates/metabolism , Animals , Anions , Kinetics , Magnetic Resonance Spectroscopy/methods , Mathematics , Models, Theoretical , Osmolar Concentration , Rabbits
16.
Biochemistry ; 30(27): 6734-41, 1991 Jul 09.
Article in English | MEDLINE | ID: mdl-2065057

ABSTRACT

The inhibitory effects of vanadium(V) were determined on the oxidation of glycerol 3-phosphate (G3P) catalyzed by glycerol-3-phosphate dehydrogenase (G3PDH), an enzyme with a thiol group in the active site. G3PDH from rabbit muscle was inhibited by vanadate, and the active inhibiting species were found to be the vanadate dimer and/or tetramer. The dimer was a sufficiently weak inhibitor at pH 7.4 with respect to G3P; the tetramer could account for all the observed inhibition. The tetramer was a competitive inhibitor with respect to G3P with a Ki of 0.12 mM. Both the dimer and tetramer were noncompetitive inhibitors at pH 7.4 with respect to NAD with Ki's of 0.36 mM and 0.67 mM. G3PDH inhibited by vanadate was reactivated when EDTA complexed the vanadate. The reactivation occurred even after extended periods of incubation of G3PDH and vanadate, suggesting that the inhibition is reversible despite the thiol group in the active site. Analogous reactivation is also observed with glyceraldehyde-3-phosphate dehydrogenase (Gly3PDH). Gly3PDH is an enzyme that previously had been reported to undergo redox chemistry with vanadate. The work described in this paper suggests vanadate will not necessarily undergo redox chemistry with enzymes containing thiol groups exposed on the surface of the protein.


Subject(s)
Glycerolphosphate Dehydrogenase/antagonists & inhibitors , Sulfhydryl Compounds/chemistry , Vanadates/metabolism , Animals , Binding Sites , Glycerolphosphate Dehydrogenase/chemistry , Glycerophosphates/metabolism , Kinetics , Magnetic Resonance Spectroscopy , Muscles/enzymology , Oxidation-Reduction , Rabbits
17.
Acta Chem Scand (Cph) ; 45(5): 456-62, 1991 May.
Article in English | MEDLINE | ID: mdl-1873111

ABSTRACT

Complex formation between tetraoxovanadate(V) and each of the nucleosides adenosine, guanosine, cytidine and uridine has been studied in a constant salt medium at pH 7. 13C- and 51V NMR studies show that only complexes with the formula V2L2 (V = vanadate, L = nucleoside) are formed, and their formation constants have been determined. They have 51V NMR resonances around -523 ppm relative to VOCl3 and they exhibit no CD in the spectral region of the charge-transfer transitions. MCD spectra were also measured, and all experiments are in accord with a molecular structure composed by two edge-sharing VO6 octahedra forming an O4V(mu-O)2VO4 skeleton with each of the nucleoside ligands bridging the two vanadium centres through the ribose 2',3'-oxygens, which are the oxygens outside the V2O6 plane. Admixture of imidazole-HCl buffer at pH 7 gives rise to additional complexes of 1:1 stoichiometry. They have been characterized by 51V NMR and CD, and their formation constants are reported. Vanadate(V) and the deoxynucleosides deoxyadenosine, deoxyguanosine, deoxycytidine and thymidine form very weak complexes which cannot be detected by 51V NMR or CD under conditions for which vanadate and the nucleosides form complexes.


Subject(s)
Nucleosides/chemistry , Vanadates/chemistry , Circular Dichroism , Magnetic Resonance Spectroscopy , Organometallic Compounds/chemistry
18.
J Biol Chem ; 266(6): 3511-7, 1991 Feb 25.
Article in English | MEDLINE | ID: mdl-1995614

ABSTRACT

Inhibition of a tartrate-resistant acid phosphatase (ACP) from Leishmania donovani and the tartrate-sensitive ACP from human seminal fluid (prostatic ACP) was examined using a series of 13 molybdate-containing heteropolyanions. The heteropolyanions were divided into four groups based on the number of molybdenum atoms they contain: Group I, Mo4; Group II, Mo6-8; Group III, Mo12; Group IV, Mo18. Two of the four groups, those consisting of compounds that contain either an Mo4 unit or an Mo18 unit with a heteroatom in the central cavity, were potent inhibitors and exhibited the highest degree of selectivity against the leishmanial and seminal fluid ACPs. The inhibition of prostatic ACP by complex E2 could be completely reversed by dialysis. Little inhibition of the acid phosphatase, beta-glucuronidase, or alpha-mannosidase from human spleen was observed with complexes B' and E2. For the seminal fluid phosphatase, the Ki values obtained with arsenate and vanadate depended markedly on pH, suggesting that, unlike most other phosphatases, the conformation of the inhibitor binding site on human seminal fluid ACP is pH-dependent. Results of competition experiments performed with various inhibitor pairs indicated that complex D2 binds to the active site of prostatic ACP while complex M binds at some site on the enzyme that affects the active site. Binding of complex M also modifies the affinity of the enzyme for other inhibitors such as vanadate. The potency of several heteropolyanion complexes and their selective inhibition of pathophysiologically significant acid phosphatases indicate that these compounds may have value as tools for study of the structure and function of this class of enzyme and perhaps in the therapy of human disease.


Subject(s)
Acid Phosphatase/antagonists & inhibitors , Leishmania donovani/enzymology , Molybdenum/pharmacology , Semen/enzymology , Animals , Anions , Humans , Hydrogen-Ion Concentration , Kinetics , Male , Tartrates
19.
Biochemistry ; 29(28): 6698-706, 1990 Jul 17.
Article in English | MEDLINE | ID: mdl-2397207

ABSTRACT

Vanadate dimer and tetramer inhibit glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides. The inhibition by a vanadate mixture containing vanadate monomer, dimer, tetramer, and pentamer was determined by measuring the rates of glucose 6-phosphate oxidation and reduction of NAD (or NADP) catalyzed by glucose-6-phosphate dehydrogenase. The inhibition by vanadate is competitive with respect to NAD or NADP and noncompetitive (a mixed type) with respect to glucose 6-phosphate (G6P) when NAD or NADP are cofactors. This inhibition pattern varies from that observed with phosphate and thus suggests vanadate interacts differently than a phosphate analogue with the enzyme. 51V NMR spectroscopy was used to directly correlate the inhibition of vanadate solutions to the vanadate dimer and/or tetramer, respectively. The activity of the vanadate oligomer varied depending on the cofactor and which substrate was being varied. The vanadate dimer was the major inhibiting species with respect to NADP. This is in contrast to the vanadate tetramer, which was the major inhibiting species with respect to G6P and with respect to NAD. The inhibition by vanadate when G6P was varied was weak. The competitive inhibition pattern with respect to NAD and NADP suggests the possibility that vanadate oligomers may also inhibit catalysis of other NAD- or NADP-requiring dehydrogenases. Significant concentrations of vanadate dimer and tetramer are only found at fairly high vanadate concentrations, so these species are not likely to represent vanadium species present under normal physiological conditions. It is however possible the vanadate dimer and/or tetramer represent toxic vanadate species.


Subject(s)
Bacterial Proteins/antagonists & inhibitors , Glucosephosphate Dehydrogenase/antagonists & inhibitors , Leuconostoc/enzymology , Vanadates/pharmacology , Glucose-6-Phosphate , Glucosephosphates/metabolism , Kinetics , Magnetic Resonance Spectroscopy , NAD/metabolism , Oxidation-Reduction
20.
Anal Biochem ; 188(1): 53-64, 1990 Jul.
Article in English | MEDLINE | ID: mdl-2221369

ABSTRACT

A kinetic method based on alkaline phosphatase has been developed to measure free trace levels of vanadium(IV) and (V). The method involves measuring the rate of the alkaline phosphatase-catalyzed hydrolysis of p-nitrophenyl phosphate with (Vi) and without (Vo) a competitive inhibitor in the assay. Michaelis-Menten kinetics for a competitive inhibitor was used to express the relationship between Vo/Vi and the inhibitor concentration. Measuring both Vo and Vi thus yields a Vo/Vi ratio that allows calculation of the competitive inhibitor concentration. Determination of free vanadium in complex fluids can be accomplished by comparing the ratio of rates of p-nitrophenyl phosphate hydrolysis with and without a sequestering agent to the ratios of rates measured on addition of a known vanadium concentration. Free vanadium(V) can conveniently be measured from 10(-7) to 10(-5) M and free vanadium(IV) can be measured at 10(-8) M and above. The error limits on the vanadium determinations range from +/- 3 to +/- 12% of the concentration under investigation depending on the conditions under which the assay was conducted.


Subject(s)
Vanadium/analysis , Alkaline Phosphatase/antagonists & inhibitors , Buffers , Hydrogen-Ion Concentration , Indicators and Reagents , Kinetics , Magnetic Resonance Spectroscopy , Radioisotopes , Reference Standards , Spectrophotometry, Ultraviolet , Temperature , Vanadium/pharmacology
SELECTION OF CITATIONS
SEARCH DETAIL
...